You are on page 1of 77

USE OF ALUMINIUM DROSS FOR SLAG TREATMENT IN SECONDARY

STEELMAKING TO DECREASE AMOUNT OF REDUCIBLE OXIDES IN


LADLE FURNACE

A THESIS SUBMITTED TO
THE GRADUATE SCHOOL OF NATURAL AND APPLIED SCIENCES
OF
MIDDLE EAST TECHNICAL UNIVERSITY

BY

ONUR AYDEMİR

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS


FOR
THE MASTER DEGREE OF MASTER OF SCIENCE
IN
METALLURGICAL AND MATERIALS ENGINEERING

JANUARY 2007
Approval of the Graduate School of Natural and Applied Sciences

─────────────
Prof. Dr. Canan ÖZGEN
Director

I certify that this thesis satisfies all the requirements as a thesis for the degree
of Master of Science.
───────────────
Prof. Dr. Tayfur ÖZTÜRK
Head of Department

This is to certify that we have read this thesis and that in our opinion it is fully
adequate, in scope and quality, as a thesis for the degree of Master of Science.

───────────────
Prof. Dr. Ahmet GEVECİ
Supervisor
Examining Committee Members

Prof. Dr. Haluk ATALA (METU,METE) ──────────────

Prof. Dr. Ümit ATALAY (METU,MINE) ──────────────

Prof. Dr. Ahmet GEVECİ (METU,METE) ──────────────

Prof. Dr. Naci SEVİNÇ (METU,METE) ──────────────

Prof. Dr. Yavuz TOPKAYA (METU,METE) ──────────────


I hereby declare that all information in this document has been obtained and
presented in accordance with academic rules and ethical conduct. I also declare
that, as required by these rules and conduct, I have fully cited and referenced
all material and results that are not original to this work.

Name, Last Name: Onur Aydemir

Signature:

iii
ABSTRACT

USE OF ALUMINIUM DROSS FOR SLAG TREATMENT IN SECONDARY


STEELMAKING TO DECREASE AMOUNT OF REDUCIBLE OXIDES IN
LADLE FURNACE

Aydemir, Onur

M.S., Department of Metallurgical and Materials Engineering


Supervisor: Prof.Dr. Ahmet Geveci

January 2007, 64 pages

In this study it was aimed to analyse refining processes such as decreasing


reducible oxide content of ladle slag with affecting parameters in low carbon
aluminum killed grades and for the research Erdemir low carbon steel grades 7112K
and 7110K are selected. There was a negative correlation between reducible oxide
amount in ladle slag and desulphurization capacity of ladle slag with metal-slag
reaction and steel internal cleanliness. To refine these properties of slag, aluminium
dross, which was aluminium production discard and has a metallic content around
%30-35 was used and after ladle treatment operation, decrease in reducible oxides
such as FetO, MnO, SiO2, P2O5 was analysed. After the study it was observed that
653 kg. of converter slag leaked during tapping of steel and SiO2 ve P2O5 content of
ladle slag had negligible change after ladle treatment. According to the results, it is
observed that initial %10-12 (FetO + MnO) content was reduced to % 4.5-5.0 (FetO
+ MnO) after ladle treatment with use of aluminium dross. Beside of this, in order to

iv
see the effect of this slag reduction on steel cleanliness, low carbon aluminium killed
grades were compared with ultra-low carbon aluminium killed grades having (FetO +
MnO) content of %16-17 in slag. It was seen that reoxidation of aluminium (loss of
dissolved aluminium) during continuous casting for ULC (ultra-low carbon) grades
is 144 ppm but for LC grades it was 94 ppm and it was being expected that ULC
steel group would have higher inclusion content after casting.

Keywords:Secondary steelmaking, slag reduction, ladle slag, reducible oxides

v
ÖZ

İKİNCİL ÇELİK YAPIMI POTA CÜRUFU MUAMELESİ İŞLEMİNDE


İNDİRGENEBİLİR OKSİTLERİN MİKTARININ AZALTILMASI İÇİN POTA
FIRININDA ALÜMİNYUM CÜRUFU KULLANIMI

Aydemir, Onur

Yüksek Lisans, Metalurji ve Malzeme Mühendisligi Bölümü


Tez Yöneticisi: Prof.Dr. Ahmet Geveci

Ocak 2007, 64 sayfa

Düşük karbonlu ve alüminyum ile deokside edilmiş çelik kalitelerinde pota


cürufunun içerisindeki indirgenebilir oksitlerin azaltılmasına yönelik uygulamaları,
etkileyen parametrelerle birlikte, inceleyen bu çalışmada Erdemir düşük karbonlu
çelik kaliteleri olan 7112K ve 7110K seçilmiştir. Pota cürufundaki indirgenebilir
oksitlerin miktarı ile, çelik içerisindeki kükürdü metal-cüruf reaksiyonuyla pota
cürufuna alma ve çeliğin iç yapı temizliği arasında ters bir ilişki vardır. Cürufun bu
özelliklerinin iyileştirilmesi amacıyla ikincil metalurji işlemleri sırasında, alüminyum
üretimi prosesi atığı olan ve %30-35 metalik alüminyum içeren alüminyum cürufu
kullanılmış ve cüruf içerisindeki ( FetO, MnO, SiO2, P2O5) gibi birincil çelik yapımı
kaynaklı olan indirgenebilir oksitlerin değişimi incelenmiştir. Çalışma sonucunda
konvertörlerden potaya sıvı çelik alınırken ortalama 653 kg. konvertör cürufunun da
potaya alındığı tespit edilmiştir. Yapılan incelemede pota cürufu içerisindeki SiO2 ve
P2O5 miktarlarının pota işlemleri sonrasında önemli miktarda değişmediği tespit
edilmiştir. Yapılan analizler sonucunda, ikincil metalurji işlemleri öncesinde cüruf
içerisinde % 10-12 olan (FetO + MnO)%, alüminyum cürufu ile yapılan işlem
sonucunda % 4.5-5.0 değerine kadar azaldığı görülmüştür. Bununla beraber cüruf
indirgeme işleminin çelik temizliğine olan etkisini görmek amacıyla, düşük karbonlu

vi
çelik grubu, üretim şekli gereği ikincil metalurji işlemi sonucunda cüruf içerisindeki
FetO ve MnO toplamı %16-17 olan ultra düşük karbonlu çelik grubu ile
karşılaştırılmış ve sürekli döküm sırasında reoksidasyona uğrayan çözünmüş
alüminyum miktarlarının (döküm sırasında azalan çözünmüş aluminyum miktarı)
ultra düşük karbonlu çelik grubu için 144 ppm ve düşük karbonlu çelik grubu için ise
94 ppm olduğu görülmüş ve bunun ultra düşük karbonlu çelik grubunun daha fazla
inklüzyon içermesine sebep olacağı beklenmiştir.

Anahtar Kelimeler: İkincil çelik yapımı, cüruf indirgeme, pota cürufu,


indirgenebilir oksitler

vii
ACKNOWLEDGMENTS

I express sincere appreciation to Prof. Dr. Ahmet Geveci for his continuous
guidance and supervision throughout the course of this thesis study.

Thanks to Recep Sarı for XRF analysis and related comments.

Erdemir Iron and Steel Works are greatly acknowledged for providing the ladle and
BOF slag analysis and performing chemical analysis.

I offer sincere thanks to my mother, my father and my brother, for encouraging and
supporting me in every stage of my academic and business carrier.

viii
TABLE OF CONTENTS

PLAGIARISM....................................................................................................... iii
ABSTRACT ......................................................................................................... iv
ÖZ ........................................................................................................................ vi
ACKNOWLEDGEMENTS ................................................................................. viii
TABLE OF CONTENTS ..................................................................................... ix
LIST OF TABLES .................................................................................................xi
LIST OF FIGURES ............................................................................................. xii
CHAPTER
1. INTRODUCTION.......................................................................................... 1
2. LITERATURE REVIEW.............................................................................. 3
2.1. Principles of Modern Steelmaking..........................................................................3
2.2.Electric-Furnace Steelmaking Processes ..................................................................5
2.3. Primary Oxygen Steelmaking Process (BOF Process)..............................................8
2.3.1. Principles ...........................................................................................................8
2.3.2. Oxidation Reactions..........................................................................................10
2.3.2.1. Oxidation of Iron .................................................................................10
2.3.2.2. Oxidation of Manganese …..................................................................13
2.3.2.3. Oxidation of Carbon .............................................................................14
2.3.2.4. Oxidation of Silicon ............................................................................15
2.3.2.5. Oxidation of Phosphorous ...................................................................16
2.3.2.6. Oxidation of Sulphur.............................................................................17
2.3.3. Reactions during Tapping ..................................................................................18
2.3.3.1. Reactions with Furnace Slag Carry-Over………….…………..…...…18
2.3.3.2. Deoxidation Reactions………………………….....………………..…19
2.4. Secondary Steelmaking Process……… ………………….……………………21

ix
2.4.1. Fundamentals of Secondary Steelmaking Process………………………….21
2.4.2. Ladle Furnace (LF)…………………………………………………….……23
2.4.3. Erdemir Steelmaking Plant Secondary Steelmaking Process………………25
2.4.4. Slags For Ladle Treatment and Refining of Steel………….………………28
2.4.4.1. Desulphurization in Ladle ……………………….…………………28
2.4.4.2. Reoxidation…………………………………………………………31
2.4.4.3. Effect of Tap Oxygen on Steel Cleanliness…………………….…..31
2.4.4.4 Effect of FeO and MnO in Slag on Steel Cleanliness………………..32
2.5. Aluminium Dross…………...……………………………………………….……37

3. EXPERIMENTAL PROCEDURE .............................................................38


3.1. Starting Materials ….……………...................................................................... .38
3.2. Outline of Experimental Procedure……………………………………….…….39
3.3. Slag Sample Preparation………………………………………….……………...40
3.4. Reducible Oxides………………………………………………..……………….41
3.5. Experimental Parameters……………………………………………………...….42

4. RESULTS AND DISCUSSION................................................................................43


4.1. Basic Oxygen Furnace Slag Chemistry .………………..................................... 43
4.2. Carry-Over Slag Weight…………………………...............................................46
4.3. Effect of Initial Slag Composition on Final Reducible Oxide %.........................50
4.4. Effect of Duration on Final Reducible Oxide % .............................................. 53
4.5. Effect of Aluminum Dross Amount on Final Reducible Oxide % ....................54
4.6. Slag Reduction by Dissolved Aluminum in Steel.................................................56
4.7. Effect of Final Reducible Oxide Content on Reoxidation of Steel......................58
5. CONCLUSIONS ...……… ………………………………………............................61
6.REFERENCES....……………… ………………………….........................................63

x
LIST OF TABLES

TABLE

2.1. Properties of Erdemir Ladle Furnace ......................................................... 27


2.2. Initial Slag Compositions used in Study, wt-%...........................................30
2.3. Effect of Ladle Slag Reduction Treatment Reported for LCAK Steel........36
3.1. Low-Carbon Aluminum Killed Steel Analysis.............................................38
3.2. Analysis of Aluminum Dross.......................................................................39
3.3. Particle Size Distribution of Al-Dross…………………………………….39
4.1. BOF Slag Analysis for LCAK Steel.............................................................44
4.2. Calculated Weight of Carry-over Slag Using Experimental Data................48
4.3. Ladle Furnace Treatment Data Including Initial and Final
Slag Compositions..............................................................................................51
4.4. Data to Show the Effect of Reaction Duration............................................53
4.5. Data to Show the Effect of Al Dross Amount (kg)....................................55
4.6. Data for Heats with and without Aluminum Dross Addition.....................56
4.7. Comparison of Low Carbon and Ultra Low Carbon Grades of Steel..........58

xi
LIST OF FIGURES

FIGURE

2.1. Flow Diagram Showing the Principal Process Steps from Raw
Material to Liquid Steel before Casting........................................................4
2.2. Sketch of Electric Arc Furnace.....................................................................5
2.3. Schematic Elevation Showing the Principal Operating Units of the
Basic Oxygen Process...................................................................................9
2.4. Effect of Slag Basicity on the Activity Coefficient of Iron Oxide in
Simple and Complex Slags ...........................................................................11
2.5. Equilibrium Ratio [ppm O] / (%FeO) related to SiO2 and P2O5..................12
2.6. Equilibrium ratio [ppm O] / (%FeO) related to Slag Basicity.....................12
2.7. Equilibrium Relation K’FeMn related to Slag Basicity...................................14
2.8. Sulphide Capacities of Slags........................................................................17
2.9. Sketch of Ladle Furnace...............................................................................23
2.10. Calculated Arc Lengths and Electric Power of Arc. ...................................24
2.11. Erdemir Steelmaking Shop Process Steps for Low Carbon Production......26
2.12. RH-KTB Type Vacuum Degasser............................................................ . 27
2.13. Effect of Unstable Oxides ( FeO + MnO) in the
Ladle Slag on Desulphurization …………………………………………….29
2.14. Influence of Different FeO Contents in Top Slag
on Desulphurization (heat size 100 tons. and
isothermal temperature 1600ºC)…………… ……………………………...30
2.15. Tap Dissolved Oxygen and Final Total Oxygen in Tundish.........................32
2.16. Relationship between FeO + MnO in Ladle Slag
and Total Oxygen in Steel after Different Ladle Practices
and in Steel in Continuous Casting Mold.....................................................33

xii
2.17. Effect of FeO + MnO Content in Ladle Slag on Lowering Dissoloved
Al (from ladle to mold) and from (ladle to tundish) and on Increasing
Alumina Content (measured in tundish).......................................................34
2.18. Correlation of FeO Content in Ladle Slag with
Ladle Slag Depth........................................................................................35
2.19. Reduction of FeO Content in Ladle Slag
by Ladle Slag Reduction Treatment............................................................36
3.1. Difference in SiO2 and P2O5 Content of Ladle Slag after Ladle Treatment...41
4.1. CaO and SiO2 Content of Erdemir BOF Slag................................................45
4.2. FeO and MnO Content of Erdemir BOF Slag................................................46
4.3. Effect of Initial Reducible Oxide Content
on Final Reducible Oxide %..........................................................................52
4.4. Effect Reaction Time on Final Reducible Oxide %......................................54
4.5. Effect Aluminum Dross Addition to Ladle Treatment Slag.........................55
4.6. Reduction (a) without and (b) with Al Dross Addition................................57
4.7. Aluminum Loss Distribution During Casting for ULC Grades....................59
4.8. Aluminum Loss Distribution During Casting for ULC Grades....................60

xiii
CHAPTER 1

INTRODUCTION

The growth in scope and extent of secondary steelmaking has been one of the
most significant features of steelmaking development in recent years and,
particularly during the past 20 years high levels of investment in new secondary
steelmaking (ladle treatment) facilities have been made by many steelmakers to
accommodate changing product needs. The reasons for this are as follows :

Primary steelmaking is mainly concerned with decarburizing in the BOF


(basic oxygen furnace) or melting in the EAF (electric arc furnace), with relatively
crude control over temperature, sulphur, and phosphorus. The complex task of fine
adjustment of composition, temperature, and steel cleanness is more effectively
carried out by secondary steelmaking which has significantly improved the technical
performance of BOF and the EAF process routes, so that the choice between them
for stringent product specifications is now made more on economic grounds and less
on technological factors.

Compared with ingot casting, continuous casting brings the benefits of higher
yield, lower rolling and energy costs, and improved quality. However, the extensive
and growing demands of continuous casting on control of analysis, temperature, and
the level of non-metallic inclusions can only be met by secondary treatment in the
ladle and it breaks the connection between the primary steelmaking operation and the
casting operation and by providing a buffer stage in the process route for logistical
control of process timing, allows them both to operate at maximum efficiency.

In general, in order to produce clean steel the melt is deoxidized and the
deoxidation products should be removed. In addition, the reoxidation by the slag
should be prevented effectively. The high oxygen potential in ladle slag may cause
contamination by reoxidation of components such as Al, Si, Ti. The oxygen potential

1
of slag is critical factor which determines the likelihood of reoxidation by slag at
ladle.
The purpose of this work was to reduce the main reducible oxides
(FetO+MnO) of ladle furnace slags by aluminium in a dross obtained from
aluminium industry which is their discard product; ladle furnace slags were made of
carry-over slag to pot during tapping of BOF furnace and additives for deoxidation
and other purposes before secondary treatment of steel.

2
CHAPTER 2

LITERATURE REVIEW

2.1 Principles of Modern Steelmaking

Practically all steel products are made at the present time by the sequence of steps
that are shown in Figure 2.1.

Iron-bearing materials containing principally iron oxides (iron ore, pellets,


sinter, etc.) are reduced to molten iron (pig iron) in the blast furnace, using the
carbon of coke as the reducing agent. In the process, the iron absorbs from 3.0 to 4.5
per cent of carbon. Iron containing 3.0 to 4.0 per cent of carbon can be used to make
iron castings (cast iron). However, most modern steels contain considerably less than
1.0 per cent of carbon, and the excess carbon must be removed from the product of
the blast furnace to convert it into steel [10].

The excess carbon is removed by controlled oxidation of mixtures of molten


pig iron and melted iron steel scrap in steelmaking furnaces to produce carbon steels
of the desired carbon content. The principal steelmaking furnaces include the basic-
oxygen furnace and electric arc furnace. Various elements like chromium,
manganese, nickel, molybdenum, etc. may be added singly or in combination to the
molten steel during or after carbon-removal process to produce alloy steels. All of
the modern steelmaking processes produce molten(liquid) metal.

3
Figure 2.1 Flow Diagram Showing the Principal Process Steps from Raw Material
to Liquid Steel before Casting [10].

After the molten steel has attained the desired chemical composition in the
steelmaking process, it is tapped (poured) from the furnace into a ladle from whence
most steel was teemed into tall usually rectangular molds where it solidifies to form
ingots. Most ingots, after being removed from the molds, were reheated to a proper
uniform temperature and rolled (in some cases, forged) into shapes known as
blooms, billets and slabs which are referred to as semi finished steel: only a relatively
small number are rolled or forged directly into finished forms. Increasing quantities
of semi finished steel are being produced by pouring liquid steel into the top of open-
bottomed molds in continuous-casting machines where it solidifies and is
continuously withdrawn from the bottom of the molds in long lengths of the desired
shapes.

4
Blooms, slabs, and billets are referred to as semi finished steel because they
form the starting material for the production by mechanical treatment (hot rolling,
cold rolling, forging, extruding, drawing, etc.) of finished steel products that include
bars, plates, structural shapes, rails, wire, tubular products, and coated and uncoated
sheet steel, all in many forms required by users of steel. Many of these products
require some form of heat treatment at the steel mill to give them the best properties
for their intended use.

The above oversimplified sequence might make it appear that the steelmaking
process is simple. Actually, the manufacture of steel in its many product forms
involves a complicated series of interrelated operations [10].

2.2 Electric-arc Furnace Steelmaking Processes

The process of making steel in basic-lined electric-arc furnace can be divided


into four periods as the melt down period, oxidizing period, the composition and
temperature adjustment period and the tapping [10] .

When charging has been completed the bank in front of the charging door is
built up with refractory material to form a dam to keep the molten metal from
slopping out the door (or doors). The door is closed and the electrodes are lowered
to about 25 mm. above the scrap. The main circuit breaker is closed, an intermediate
voltage is selected with proper current setting on the rheostats, and the arcs are struck
under automatic control. After 1 to 3 minutes (to allow the electrodes to bore into the
scrap), maximum voltage and current should be applied for the fastest possible
melting of scrap. The initial slow start is to shield the lining and roof from the heat of
the arc.

5
The melting period in the basic electric furnace is the most expensive period in
its operation because power and electrode consumption are at the highest rate during
this interval.

Figure 2.2 Sketch of Electric Arc Furnace

The electrode melts the portion of the charge directly underneath and around
them, and continue to bore through the metallic charge, forming a pool of molten
metal on the hearth. From the time the electrodes bore through the scrap and form a
pool of molten metal on the hearth, the charge is melted from the bottom up by
radiation from the pool, by heat from the arc, and by the resistance offered to the

6
current by the scrap. Burnt lime is usually blown on top of the bath near the end of
the melting period of the first scrap charge. If blowing facilities are not available,
burnt lime is added usually to the furnace before the second scrap charge is made.
Heating of the charge material is continued until it is completely melted.

Oxidation occurs in varying degrees from the time the molten metal begins to
form until the entire charge is in solution. During this period, phosphorus, silicon,
manganese, carbon, and iron are oxidized. Oxygen for these, as well as for other
oxidizing reactions is obtained from:

• Oxygen gas injected into the bath


• Oxygen in the furnace atmosphere
• Calcination of limestone (only small amounts of limestone are used on a flat
low-carbon steel bath to minimize overheating of the furnace roof by heat
radiation from a smooth bath)
• Oxides of alloying elements added in the furnace
• Ore, cinder and scale

The oxidizing materials added to the bath react with carbon in the bath to form
carbon monoxide. If too much of the oxidizing material is added at one time, the
carbon monoxide formation may be vigorous enough to eject slag and steel out of the
furnace. The direct use of oxygen is extremely important in modern practice from the
standpoint of rapidly removing carbon from the bath.

Enough carbon should be present in the charge so that the carbon content of
the steel, when melted, is 0.15 to 0.25 % higher than the desired tap carbon content.
This excess carbon is removed with oxygen and forms carbon monoxide gas, which
bubbles out of the steel. This bubbling is called the carbon boil, which stirs the bath
and makes it more uniform in composition and temperature. This facilitates meeting
steel composition and temperature specifications. The carbon boil also removes some
of the hydrogen and nitrogen from the steel which is generally desirable.

7
During the oxidation period, the reactions that occur in the bath of the basic
electric arc furnace are similar to those in the basic open-hearth furnace and the basic
oxygen furnace, except that the electric furnace bath can be made hotter. Hence,
there is more chance of phosphorus reversion unless the slag is strongly basic.

Most carbon steel and low-alloy steel grades made in electric-arc furnaces are
made by a single slag process. This means that the slag that forms during melting of
the charge is not replaced by another slag as in the double slag process. In the single
slag process, steelmaking is finished by adjusting the composition and temperature of
steel bath to the desired values, after which the steel is tapped from furnace into a
ladle. As soon as a charge is melted, a steel sample is taken and analyzed in the
chemical laboratory. From this analysis, the required steel composition adjustments
are determined. Several temperature measurements of the steel are made to
determine the temperature adjustment required.

In tapping a heat, the electrodes are raised high enough to clear the bath in a
tilted position, the tap hole is opened, and the furnace is tilted by a control
mechanism so that the steel is drained from the furnace into a ladle. The ladle is
usually held by a teeming crane during tapping to minimize exposure of the stream to
air and minimize erosion of ladle refractory. The slag may be tapped before, with, or
after the steel depending on the particular operation.

2.3 Primary Oxygen Steelmaking Process (BOF Process)

2.3.1 Principles

A schematic elevation showing the principal operating units of the basic


oxygen process is shown in Figure 2.3 [9].

Most commonly, basic oxygen furnace operation starts with charging of


furnace with the proper amount of scrap and molten iron typically containing 4.1

8
%C, 0.35 %Mn, 0.1 %P, 0.5 % Si. It is turned upright and the oxygen lance is
lowered to a predetermined position above the surface of the bath. Pure oxygen gas
issues from the jet nozzle at high velocity under a pressure that is normally held
between 14 and 16 bars. The action of the oxygen jet is partly chemical and partly
physical. The oxygen immediately starts reacting with the silicon, generating heat
and producing molten silica. Slag-forming fluxes (chiefly burnt lime, dolomitic lime,
and fluorspar) are added in controlled amounts from an overhead storage system
immediately after oxygen ignition. These materials, which serve to produce a slag of
the desired basicity and fluidity, are added through an inclined chute built into the
side of a water-cooled hood that covers the furnace.

Figure 2.3. Schematic Elevation Showing the Principal Operating Units of the Basic
Oxygen Process [9]

9
As the blow progresses, an emulsion is formed with the slag and metal, and
the other metalloids are oxidized. Carbon monoxide is evolved, which gives rise to a
vigorous boiling action and accelerates the refining metallurgical reactions.

When the blow is completed, as determined by the furnace operator utilizing


the results of calculations by model computer, the lance is withdrawn and the furnace
is tilted to a horizontal position toward the charging side. A temperature reading is
secured with an immersion-type thermocouple, and a sample of the steel is obtained
and sent to the nearby chemical laboratory for analysis. This analysis determines the
individual chemical elements of the steel.

2.3.2 Oxidation Reactions

2.3.2.1 Oxidation of Iron

With hypothetical pure liquid FeO as the standard state, the activity of iron
oxide is derived from the concentration of dissolved oxygen in liquid iron that is in
equilibrium with the slag. For the reaction equilibrium [9]

FeO(l) = Fe + [O] (1)

KO = [aO] / aFeO (2)

where aO = [%O]γO, using log γO = -0.1 * [%O]. The temperature dependence of the
equilibrium constant KO, as derived from the standard free energy equation for
reaction (1), is given below:

log KO = -5730/T + 2.397 (3)

10
A wide variety of formulations is carried out to represent the composition
dependence of the iron oxide activity or activity coefficient in complex slags. Within
the limits of uncertainty of the experimental data on slag-metal reaction equilibrium,
there is a decisive correlation between the activity coefficient of FeO and the slag
basicity B as shown in Figure 2.4. It is evident that γFeo reaches a peak at basicity of
about B = 1.8 .

Figure 2.4. Effect of Slag Basicity on the Activity Coefficient of Iron Oxide in
Simple and Complex Slags [9]

In steelmaking slags, the total number of g-mols of oxides per 100 g of slag is
within the range 1.65 ± 0.05. Therefore, the analysis of the slag-metal equilibrium
data, in terms of the activity and mol fraction of iron oxide, can be transposed to a
simple relation between the mass ratio [ppm O] / (%FeO) and the sum of the acidic
oxides %SiO2 +0.84 x %P2O5 as shown in Figure 2.5. For basicities of B>2, this
mass ratio as function of slag basicity is given in Figure 2.6 [ 9 ].

11
Figure 2.5. Equilibrium Ratio [ppm O] / (%FeO) Related to SiO2 and P2O5 [9]

Figure 2.6. Equilibrium Ratio [ppm O] / (%FeO) Related to Slag Basicity [9]

12
2.3.2.2 Oxidation of Manganese

The slag metal distribution ratio is governed by two interrelated reaction


equilibria. One reaction is oxidation of Mn with FeO in slag [9].

(FeO) + [Mn] = (MnO) + [Fe] (4)

KFeMn = aMnO / aFeO * [%Mn] (5)

where the oxide activities are with respect to pure liquid oxides.

The second reaction to be considered is

[Mn] + [O] = (MnO) (6)

KMn = aMnO / [%Mn] [aO] (7)

and temperature dependence of these equilibrium constants KFeMn and KMn are,

log KFeMn = 7452/T – 3.478 (8)

log KMn = 13182/T –5.875 (9)

For the FeO and MnO exchange reaction involving the oxidation of
manganese in steel, the equilibrium relation may be described in terms of the mass
concentrations of oxides.

K’FeMn = (%MnO) / (%FeO) [%Mn] (10)

13
where, the equilibrium relation K’FeMn depends on temperature and slag composition.
The values of K’FeMn against the slag basicity for tapping temperature from 1600°C
to 1700°C is plotted in Figure 2.7. In most of the practices tapping temperature is
around 1650°C for which the equilibrium K’FeMn is about 1.8 ± 0.2.

Figure 2.7 Equilibrium relation K’FeMn related to slag basicity [9]

2.3.2.3 Oxidation of Carbon

For the reaction,

CO(g) = [C] + [O] (11)

the equilibrium constant for low alloy steels containing less than 1% C for
steelmaking temperatures, is

K = [%C][ppmO] / pCO(atm) = 20 (12)

14
With respect to slag-metal reaction, the equilibrium relation for carbon oxidation will
be,

(FeO) + [C] = CO(g) + [Fe] (13)

KFC = pCO (atm) / [%C] aFeO (14)

log KFC = - 5730/T +5.096 (15)

For the reaction,

(MnO) + [C] = CO(g) + [Mn] (16)

the following equilibrium relations apply

KMC = pCO (atm) [%Mn] / [%C] aMnO (17)

log KMC = -13182/T + 8.574 (18)

2.3.2.4 Oxidation of Silicon

For low silicon contents mass concentrations of Si and O in the equilibrium


constant for the reaction are being used [9] .

[Si] + 2[O] = (SiO2) (19)

KSi = aSiO2 / [%Si][%O]2 (20)

where the silica activity is with respect to solid SiO2 as the standard state. The
temperature dependence of KSi is given by

15
log KSi = 30410/T – 11.59 (21)

For the BOF tap temperatures KSi ≈ 3 x 104 and for lime-saturated slags aSiO2
< 0.01, therefore for 800 ppm of O in steel at turndown the equilibrium content of
silicon in the steel would be less than 1 ppm Si. In practice the steel contains 0.003 to
0.005% Si at turndown. Evidently, at these low concentrations of silicon the reaction
kinetics are no longer favourable to reach slag-metal equilibrium with respect to
silicon oxidation.

2.3.2.5 Oxidation of Phosphorus


Formulation of the phosphorus reaction is [9]

[P] + 5/2[O] + 3/2(O2-) = (PO43- ) (22)

At low concentration of [P] and [O], as in most of the experimental melts,


their activity coefficients are close to unity, therefore mass concentrations can be
used in formulating the equilibrium relation KPO for the above reaction.

KPO = (%P)[%O]-2.5 / [%P] (23)

The equilibrium relation KPO known as the phosphate capacity of the slag,
depends on temperature and slag composition. After several studies [9] it is
understood that sum of basic oxides is considered to be key parameter in describing
the composition dependence of the phosphate capacity of slag.

BO = %CaO +%CaF2 + 0.3 * %MgO (24)

The temperature and slag composition dependent formula of phosphate capacity of


slag is

log KPO = 21740/T – 9.87 + 0.071 * BO (25)

16
2.3.2.6 Oxidation of Sulphur

The sulphur transfer from metal to slag is reduction process as represented by


the equation

[S] + (O2-) = (S2-) + [O] (26)

for which the state of slag-metal equilibrium is,


KSO = (%S) [%O]/ [%S] (27)

The concentration of acidic oxides, e.g. %SiO2 + 0.84 * %P2O5 is to be a better


representation of the dependence of KSO on the slag composition. See Figure 2.8.

Figure 2.8 Sulphide Capacities of Slags [9]

17
In view of the relationship between the ratio [ppm O] / (%FeO) and the sum of
the acidic oxides, the sulphide capacity of the slag may be represented also by
formula given below.

KS = (%S)(%FeO) / [%S] (28)

2.3.3 Reactions During Tapping

2.3.3.1 Reactions with Furnace Carry-over Slag

Some of the additions to steel as Fe-Mn, Fe-Si, Al, etc are done during taping
of BOF furnace, before ladle treatment (secondary metallurgy). The percentages of
utilization of the ladle additions for low carbon heats are 85-95% for Mn, 60-70% for
Si, 35-65% for Al. Some of manganese is lost during ladle additions because of the
vaporization of metal. Part of slag from BOF furnace is carried over to the ladle for
steel. Losses in the ladle additions of aluminium and silicon are due to reactions with
this furnace slag that is carried into the tap ladle. There is also the iron oxide rich
skull accumulating at the converter mouth, some of which falls into the ladle during
furnace tapping and reacts with aluminium and silicon.
The reaction of aluminium and silicon with the furnace slag and fallen
converter skull is represented in a general form by the following reaction

Fe(Mn)Ot + Al(Si) = Fe(Mn) + Al(Si)Ot (29)

According to Turkdogan [9] using the average molecular masses and


assuming 80% Fe3O4 for the converter skull, the following approximate relation is
derived for the loss of aluminium and silicon to the ladle slag for 220 tons steel in the
tap ladle.

[%Al + %Si]s = ∆(%FeO + %MnO)Wfs x 10-6 + Wsk x 10-4 (30)

18
where,
Wfs = mass of furnace slag carryover, kg
Wsk = mass of fallen converter skull, kg
∆(%FeO + %MnO) = decrease in oxide contents of furnace slag during tapping.
[%Al + %Si]s = loss of Al and Si to the ladle slag, %

Another consequence of slag carryover is phosphorus reversion to the Al-killed


steel in the tap ladle. For a 220 tons heat, the phosphorus reversion ∆[ppm P] is
represented by:

∆[ppm P] = 0.045 ∆(%P) Wfs (31)

where ∆(%P) is the decrease in the phosphorus content of the furnace slag carried
into the tap ladle.

For high phosphorus reversion of 45 ± 10 ppm., the mass of slag carry-over is


estimated to be Wfs = 3214 ± 714 kg [9]. In most cases of furnace tapping, the
phosphorus reversion is about 20 ± 10 ppm for which the estimated slag carry-over is
about 1430 ± 714 kg.

2.3.3.2 Deoxidation Reactions

Deoxidation reactions can be described using the deoxidation equilibrium


constant (K). The deoxidation reaction, when the alloying element (M) is added to
the steel, can be represented by :

MxOy = xM + yO (32)

The deoxidation constant assuming pure MxOy forms (i.e. unit activity for
MxOy) is given by :

19
K = (hM)x (hO)y (33)

where hM and hO are the Henreian activities defined such that activity of the
components is equal to its weight percent at infinite dilution in iron. [13]

hi = fi (%i) (34)

where
fi = activity coefficient of i
%i = weight percentage of i

For most low alloy steels encountered in ladle metallurgy the activity
coefficient can be taken as unity and equation (33) reduces to :

KM = (%M)x (%O)y (35)

There are primarily three elements used in steel deoxidation to remove


dissolved oxygen while tapping from converter to ladle. Manganese (as low and high
C ferro-alloy), silicon (or as silicomanganese alloy) and aluminium (about 95-98%
purity).

For low carbon aluminium killed steel grades only aluminium bars are used
to remove oxygen. The minimum amount of aluminium required to deoxidize the
steel is computed based on the following reaction [6]:

2 [Al] + 3[O] = Al2O3 (s) (36)

For the reaction,


Al2O3(s) = 2[Al] + 3[O] (37)

20
K= [%Al]2[ppm O * fo]3 / aAl2O3 (38)

log K = - 62680/T + 31.85 (39)

In practice aluminium is used in conjunction with CaO and a complex


aluminate will form; either CA, CA2 or C12A7 (A represents Al2O3 and C represents
CaO). It is desirable to form a liquid calcium aluminate nearly saturated with CaO so
that the resulting inclusion is C12A7 in the solidified steel. For example, for 0.01% Al
the activity of oxygen will be less than 2 ppm, whereas, for Al alone it would be
about 10 ppm.
A second advantage of Al-CaO deoxidation is that, a liquid calcium
aluminate inclusion forms which floats out more readily than Al2O3 and liquid
inclusions will not clog casting nozzles.

(CaO) + 2[Al] + [3O] = (CaO.Al2O3) (40)

2.4 Secondary Steelmaking Process

2.4.1 Fundamentals of Secondary Steelmaking Process

The purpose of “Secondary Steelmaking” is to produce clean steel, steel


which satisfies stringent requirements of surface, internal and micro cleanliness and
requirements of mechanical properties. Ladle metallurgy (or secondary metallurgy)
is a secondary step of the steelmaking process often performed in ladle after the
initial refining process in a primary furnace is completed.

Although satisfactory for making steels for most of the steel applications,
conventional steelmaking and refining practices such as BOP, Q-BOP and electric
arc-furnaces could not consistently achieve the specifications the special steels had to
meet. To remain competitive and maintain production, steelmakers reluctantly
accepted the secondary steel refining processes. It is found that they were able to

21
exercise a control over the many processing conditions which contributed to an even
higher quality steel including [9] :

• Teeming temperature, especially for continuous casting operations


• Deoxidation
• Decarburization (ease of production of steels with low carbon levels of less
than 0.03%)
• Additional adjustment for chemical composition
• Increasing production rates by decreasing refining times in the BOF

Secondary steelmaking processes are adopted primarily to achieve various


objectives. The fulfilment of these objectives result in a steel which meets the desired
stringent requirements. These objectives (may be called as goals of secondary
steelmaking) include:

• Control of gases (decreasing the concentration of oxygen and hydrogen in


steel)
• Low sulphur contents (less than 0.01% and to as low as 0.002% )
• Micro cleanliness (removal of undesirable non-metallic, primarily oxides and
sulphides)
• Inclusion morphology (Since steelmakers cannot remove undesirable oxides
completely, this step allows steelmakers to change the composition and/or
shape of the undesired matter left in the steel to make it compatible with the
mechanical properties of the finished steel.)
• Mechanical Properties (Charpy V-notch toughness, transverse properties, and
ductility in through-thickness of plates)

22
2.4.2 Ladle Furnace (LF)

The ever increasing users demand for higher quality steel particularly for
critical applications and the advent of sequential continuous casting, made it
necessary for steel plants to install ladle furnaces for steel reheating that is needed for
extended time for steel refining and the adjustment of steel temperature for the caster
on a correct timely basis.

There are several types of ladle furnaces for arc reheating of steel designed by
different manufacturers but the basic features of the ladle furnace is given in Figure
2.9 [10].

Figure 2.9. Sketch of Ladle Furnace [9]

The following statements summarise the basic requirements for efficient arc
reheating in the ladle furnace.

23
1. To shorten reheating (submerged arcing) time, minimise refractory erosion
and increase the efficiency of energy consumption a large capacity
transformer and argon stirring with bottom porous plug at a flow rate of about
0.33 m3/min, is used .
2. For high heating efficiency with a stable submerged arc and low refractory
erosion, the thickness of the slag layer should be about 1.3 * arc length. The
calculated arc length and electric power of arc are shown Figure 2.10 [9] .
3. With the tap voltage of 385 to 435 V and secondary current 40 to 50 kA the
liquid steel temperature is raised at the rate of 3.5 to 4.5 °C /min.

After arc reheating and trim additions for composition adjustment, the rate of
argon bubbling should be reduced to a low level of 0.10 m3 / min. to float out the
residual deoxidation and reoxidation products known as inclusions.

Figure 2.10 Calculated arc lengths and electric power of arc [9] .

24
2.4.3 Erdemir Steelmaking Plant Secondary Steelmaking Process

In Erdemir Steelmaking Plant there exists three Basic Oxygen Furnaces


(BOF), two ladle furnaces and one recirculating type (RH-KTB) vacuum chamber.
The process steps for low carbon steel are shown in Figure 2.11. The vacuum
degasser is used to produce ultra low carbon steel grades and it is shown in Figure
2.12.

25
26
Figure 2.11 Erdemir Steelmaking Process Steps for Low Carbon Steel Production
Vacuum
Pump Oxygen Blowing
Lance

Vacuum
Chamber Reaction
Surface

Ladle
Snorkels

Figure 2.12 RH-KTB Type Vacuum Degasser

Typical properties of the ladle furnace are given in Table 2.1.

Table 2.1 Properties of Erdemir Ladle Furnace

Power 18 MVA
Primary Voltage 13.8 KV
Secondary Voltage 331 V
Secondary Current 38 KA
Electrode Diameter 500 mm.
Heating Rate 3-5 ° C / min

27
2.4.4 Slags for Ladle Treatment and Refining of Steel

Slags of certain type play an important role in the ladle treatment of steel,
particularly in the production of killed steel low in residual concentration of oxygen
and sulphur. The deoxidation of steel with aluminium to low levels of oxygen can be
greatly enhanced by adding burnt lime together with aluminium so that the reaction
product would be lime saturated liquid calcium aluminate. The fluxing of the
reaction product alumina with lime is facilitated by also adding some fluorspar
together with lime. Such a ladle treatment also brings about some removal of sulphur
from the steel. In fact, lime and aluminium are considered to be the two essential
ingredients of the ladle additions for effective deoxidation and desulphurization of
the steel [8].

Also since steel refining and continuous casting operations control steel
cleanliness, inclusion removal capacity of slag is important in terms of these
parameters. Systematic studies show that ladle treatment lower inclusions by 65-75
%, the tundish removes inclusions by 20-25%, and, although some reoxidation
occurs the mould removes 5-10 % of the inclusions. Thus ladle operations are
particularly important and include control of tap oxygen, reducible oxide amount
(FetO and MnO ) in the slag and ladle stirring [8].

2.4.4.1 Desulphurization in the Ladle Furnace

The basic chemical reaction describing the thermodynamics of


desulphurization with synthetic slags can be written as:

3(CaO) +2[Al] +3[S] = 3(CaS) + (Al2O3) (41)

The Al and S are dissolved in the metal and the CaO, Al2O3 and CaS are in
the slag. From thermodynamics it is shown that [13] the quantity (S) / [S] [Al]2/3 is

28
function of slag composition only and the slag that gives the highest value of this
parameter is the optimum slag for the process. In ladle furnace tests it is found that
[13] the best slag is one containing 10%SiO2, 30%Al2O3, and 60%CaO. CaF2
additions are beneficial for several reasons. It increases the sulphide capacity of the
slag, the fluidity of slag, and the amount of CaO that can be dissolved in the liquid
slag.
Also presence of unstable oxides (FeO+MnO) in ladle slag can reduce
desulphurization capacity of slags significantly as it is shown in Figure 2.13.

Figure 2.13 Effect of Unstable Oxides ( FeO + MnO) in the Ladle Slag on
Desulphurization

In order to evaluate the influence of the FeO content in the slag on


desulphurization and aluminium loss, a parameter study is made by Andersson et al.
[1] for three different slag compositions where the initial FeO contents in the slag
were 0, 2 and 6%. The different initial slag compositions in the study are specified in
Table 2.2.

29
Table 2.2 Initial Slag Compositions (wt%) Used in the Study by Andersson et al. [1]

Figure 2.14 shows the influence of different initial contents of FeO in the top
slag on desulphurization at an isothermal temperature of 1600ºC. At 0% FeO, the
average sulphur content decreased from 0.023 to 0.013% within 10 min. When the
initial FeO content increased, the desulphurization rate decreased. At 2% initial FeO,
the average sulphur content was 0.015% after 10 min treatment. When the initial FeO
content was increased even further to 6%, the average sulphur content was 0.020 %
after 10 min.

Figure 2.14. Influence of Different FeO Contents in Top Slag on Desulphurization


(heat size 100 tons., isothermal temperature 1600ºC) [1]

30
2.4.4.2. Reoxidation

Once the steel has been properly deoxidized, care should be taken to avoid
reoxidation. There are many problems concerning steel cleanliness and castability
caused by reoxidation.
Reoxidation increases; the oxygen content of the steel bath, contributes
towards new groups of oxide inclusions, reduces deoxidant efficiency and increases
the opportunity for nozzle blockage at the caster. It can occur both during and after
deoxidation and the main oxygen sources are; air entrainment, oxygen containing
ladle slag, refractory lining and ladle glaze or bottom slag that results from earlier
heats in the ladle. It can be particularly prevalent during intensive ladle stirring when
a slag-free eye can open-up and expose the steel surface directly to the atmosphere.
[ 7,12,13]
When the ladle top slag contains high level of reducible oxides (particularly
FeO and MnO), then oxygen transfer from slag to steel can be a principal source of
reoxidation, especially during ladle stirring. This situation often arises when a
significant amount of tapping slag is carried over into the ladle. [ 4,5]

2.4.4.3. Effect of Tap Oxygen on Steel Cleanliness

Tap oxygen content is measured during tapping the molten steel in ladle from
BOF or before deoxidant addition. The tap oxygen content is typically high, ranging
from 450–1000 ppm for most of the basic oxygen furnaces. Aluminium additions to
deoxidize the melt create large amounts of Al2O3 inclusions. This suggests that a
limitation on tap oxygen content should be imposed for clean steel grades. However,
as shown in Fig. 2.15 [8], there is no correlation between tap oxygen and steel
cleanliness (given here as final total oxygen content of steel in tundish). This is
consistent with claims that 85% of the alumina clusters formed after large aluminium
additions readily float out to the ladle slag, and that the remaining clusters are
smaller than 30µm. Naturally, the decision to ignore tap oxygen depends on the time

31
available to float inclusions and on the availability of ladle refining, which can
remove most of the generated inclusions.

Figure 2.15. Tap Dissolved Oxygen and Final Total Oxygen in Tundish [8]

2.4.4.4.Effect of FeO and MnO in Slag on Steel Cleanliness

An important source of reoxidation is the carryover of slag from the


steelmaking furnace to the ladle, which contains a high content of FeO and MnO [8].

These molten oxides react with the dissolved aluminium to generate solid
alumina in the liquid steel, owing to the strong favourable thermodynamics of the
following reactions:

3(FeO) + 2[Al] = (Al2O3) + 3[Fe] (42)

∆G° = - 853 700 + 239.9T (J mol-1) (43)

3(MnO) + 2[Al] = (Al2O3) + 3[Mn] (44)

32
∆G° = -337 700 + 1.4T (J mol-1) (45)

The higher is the FeO and MnO content in the ladle slag, the greater is the
potential for reoxidation and the corresponding generation of alumina inclusions.
Many slivers in the final product have been traced to reoxidation that originated from
FeO in the ladle slag [8,11]. Figure 2.16 [8] shows how total oxide in the ladle,
tundish and mould correlates roughly with the %FeO + %MnO in the ladle slag.

Figure 2.16. Relationship between FeO + MnO in Ladle Slag and Total Oxygen in
Steel after Different Ladle Practices and in Steel in Continuous Casting Mould [8].

These slag impurities are particularly damaging when the ladle lining material
is high in silica. Figure 2.17 [8] shows a similar influence on the loss of dissolved Al
and the increase of tundish inclusions.

33
Figure 2.17 Effect of FeO + MnO Content in Ladle Slag on Increasing Dissolved Al
and so on Increasing Alumina Content of Steel [8]

Countermeasures to lower FeO and MnO contamination are summarized as


follows:

1) Minimize slag carryover from steelmaking furnace to ladle during tapping:

• Increasing aimed carbon at the end of oxygen blowing, avoiding reblows,


thus minimizing the dissolved oxygen content in the steel, can lower the
amount of FeO in the furnace slag.
• Use of a sublance in the BOF substantially lowers the frequency of reblows.
• An efficient slag stopper, such as a slag ball (that floats in steel and sinks in
slag), can lower the amount of furnace slag carried over to the ladle during
tapping to 3 kg/t steel. To improve yield, other sensors are available, which
detect the slag after a little has been carried over. These include infra-red and
electromagnetic systems.
• The furnace geometry and tap hole location can be designed to minimize
vortexing and slag carryover.
• A thick ladle slag layer after tapping suggests high slag carryover problems,
as shown in the rough relationship in Fig. 2.18. Ladle slag depth after
skimming should be controlled to 25–40 mm.

34
Figure 2.18 Correlation of FeO content in Ladle Slag with Ladle Slag Depth [8]

2) Use a ladle slag reduction treatment.

Another way to lower the FeO + MnO content of the ladle slag is to add a
slag conditioner (i.e. slag reduction or deoxidation treatment), which is a mixture of
aluminium and burnt lime or limestone. Table 2.3 [8] summarizes the decrease in
FeO + MnO content after ladle slag reduction treatment at several steel plants. On the
average, this treatment lowers FeO + MnO to below 5 %, as shown in Fig. 2.19.
Minimizing slag carryover, together with adding a basic ladle slag and using a basic
lining, was found to lower the ladle slag FeO + MnO content to less than 1–2%,
which lowered the total oxygen in steel to 10 ppm for Low Carbon Aluminium
Killed (LCAK) Steel.

35
Table 2.3 Effect of Ladle Slag Reduction Treatment Reported for LCAK Steel [8]

FeO+MnO composition in ladle slag


Steel Works Before reduction After reduction Year
treatment treatment
FeO 3.9%, MnO 1.6 FeO FeO 1.6%, MnO 0.9 1993
LTV Steel, Cleveland Works
25.9%, MnO 2.9% FeO 4.2%, MnO 2.0% 1993

Inland Steel, No.4 BOF Shop FeO 8.1%, MnO 5.2% FeO 2.4%, MnO 1.4% 1990
National Steel, Great Lake Div. FeO 25% FeO 8%, Best 2.0 % 1994
USS/Kobe Steel Company FeO 30% FeO 1.23%, Best 0.64% 1991
Algoma Steel, Canada FeO 1.5%, MnO 0.8% 1999
Sollac Dunkirk, France FeO 12-25% FeO 2-5% 1997
Krupp Stahl AG, Bochum Steelwork FeO + MnO %<1 1991
POSCO, Kwangyang Works, Korea FeO + MnO 9-18% FeO + MnO 3-5% 1998
Kawasaki Steel, Mizushima Works FeO 2% 1996
Chian Steel, Taiwan FeO 26.8%, MnO 4.7% FeO 6.8%, MnO 5.5% 1996
WISCO No.3 Works, China FeO <1% 2000

Figure 2.19. Reduction of FeO Content in Ladle Slag by Ladle Slag Reduction
Treatment [8].

36
2.5 Aluminium Dross

Two methods are applied in the industry to manufacture aluminium products.


The first one depends on the bauxite and cryolite as the main raw materials, where
pure alumina (Al2O3) is extracted by Bayer process. The second method consists of
refining, remelting, and alloying the scrap of the manufactured aluminium [14]. This
process is applied in “Akdağ Aluminium” and “Şahinler Metal” company. The
factory sells the coarse fraction of the dross to local enterprises which manufacture
low grade aluminium wares.
Aluminium dross is formed at the surface of the molten metal due to its
uncontrolled reaction with the furnace atmosphere. It may contain up to 70% free
aluminium in an oxide layer. Many trials have been done all over the world to
minimize the formation of dross and to recycle it in most economic ways. The
chemical and phase composition of the Al-dross vary within a wide range according
to the technology used for Al-metal production and recycling. The major phases
detected are aluminium oxides with different crystalline forms and metallic
aluminium [14,15].

37
CHAPTER 3

EXPERIMENTAL PROCEDURE

3.1. Starting Materials

As stated before the purpose of this thesis work was to decrease the amount of
reducible oxides (FetO+MnO) in the carry-over slag (it can not be prevented to leak
during tapping of steel in the ladles) from BOF furnace by adding some aluminium
dross (slag conditioner) in the ladle furnace. These reducible oxides in the slag
increased the oxidation potential and so oxide content of steel. Some analytical plant
data from production of Ereğli Iron and Steel Works Co. (Erdemir), were specially
followed and taken for this purpose. The analytical data taken are for slags in
connection with Erdemir’s special low carbon aluminium killed steel grades, 7112K
and 7110K; the chemical analysis of these steels are given in Table 3.1.

Table 3.1. Low-Carbon Aluminium Killed Steel Analysis*

Grade %C % Mn %P %S %Al
7110K 0.02-0.05 0.15-0.25 0.020 max. 0.020 max. 0.020-0.060
7112K 0.02-0.04 0.10-0.20 0.015 max. 0.015 max. 0.020-0.060
*Rest of analysis can be taken as Fe, excluding Oxygen in ppm levels.

The aluminium dross is procured from Turkish Akdağ Aluminium and


Şahinler Metal companies. The average composition ranges and particle size
distribution of aluminium dross used are shown in Tables 3.2 and 3.3, respectively.

38
Table 3.2. Analysis of Al Dross

%Al2O3 %Al %SiO2 %Fe %C %K %Mg %Cu

55-60 30-35 5-10 0.5-1.0 0.10-0.20 0.5-1.0 0.5-1.0 0.10-0.20

Table 3.3 Particle Size Distribution of Al-Dross

Size <0.5 mm. 0.5-30 mm >30 mm


Weight % 4.8 91.1 4.1

3.2 Outline of Experimental Procedure

In Erdemir Steelmaking Shop, primary steelmaking starts with oxygen


blowing in basic oxygen furnaces (converters). For characterizing of converter slag,
while taking steel samples for usual process practice, slag samples are also taken
with a 2.5 m steel rod from the contact surface of liquid steel with slag just after the
oxygen blowing.

For the production of low-carbon aluminium killed steel usual practice followed
the sequence given below, after steel composition was determined:

• Steel was tapped from converter to a ladle and during tapping period (4-6
minutes) aluminium bar addition (changes with dissolved oxygen amount ) to
kill the steel and 600 kg of burnt lime addition was done to the escaped
(carry-over) converter slag.

39
• Ladle with 120 tons of liquid steel was brought to ladle furnace station with
crane. Before taking initial steel and slag samples, steel was heated for 3
min. by arc-heating and bottom-stirred with argon (0.33 m3/min).

• After this pre-treatment process, steel samples with a spoon, and slag samples
with a steel rod from contact surface, were taken.

• After obtaining initial steel composition, determined by spectometer instantly


, according to aimed time for sending treated steel to continuous casting,
heating and alloying was done at the ladle furnace to set the chemical
composition of steel in a narrow specification range. Just before sending the
ladle from ladle treatment station to continuous casting plant, final steel and
slag samples were taken in the same way initial samples were taken.

3.3 Slag Sample Preparation

As it was explained in part 3.2 slag samples were collected with a steel rod
and prepared for analysing in XRF Spectrometer. Slag sample preparing procedure
was made up of the following steps:

• Representative slag pieces (3-4 cm.) were collected from the surface of the
steel rod.
• These slag samples were crushed and ground with a grinding mill to 100
microns.
• Metallic iron which remained suspended in slag was seperated with a
magnet.
• Ground slag particles were pressed into cylindrical pellets in a hydraulic press
under 100 kN.

40
3.4. Reducible Oxides

As it was explained in literature review section, FetO and MnO were the main
reducible oxides in ladle slag which affected adversely the desulphurization capacity
of slag and internal cleanliness of steel. Besides these oxides, SiO2 and P2O5 which
formed during primary oxygen steelmaking process, also leaked with carry-over slag
from converters during tapping; these can be also reduced by dissolved aluminium
during ladle treatment. In Erdemir Steelmaking Plant, there were two alternative
converter slag stoppage systems (magnetic wave system and thermal camera) and
amount of carry-over slag which escaped to ladle was kept between 264 kg. and 1143
kg. . For the carry-over slag with such an amount, reduction of SiO2 and P2O5 were
also determined according to initial and final ladle slag analysis. For 7112K and
7110K steel grades, average initial SiO2 and P2O5 contents of ladle slag were 4.96%
and 0.37%, respectively. For 80 trial heats, differences in P2O5 and SiO2 percentages
during ladle treatment are shown in Figure 3.1.

16
25,0
14

20,0 12

10
15,0
%of Heats
% of Heats

10,0 6

4
5,0
2

0
0,0
-1 -0,8 -0,6 -0,4 -0,2 0 0,2 0,4 0,6 0,8 1 More 0 0,04 0,08 0,12 0,16 0,2 0,24 0,28 0,32 0,36 0,4 More
Difference (% ) in P2O5 After Ladle Treatment
Difference (% ) in SiO2 After Ladle Treatment

Figure 3.1 Difference in SiO2 and P2O5 content of ladle slag after ladle treatment

According to these results, during secondary steelmaking process, in low


carbon aluminium killed steel grades, on the average 0.10% of SiO2 and 0.17% of
P2O5 were reduced by aluminium. Therefore, under Erdemir Steelmaking Plant

41
process conditions, reduction of SiO2 and P2O5 could be neglected and for this study
FetO and MnO were taken into account as reducible oxides.

3.4 Experimental Parameters

Parameters affecting the final reducible oxide amount of the ladle slag were
determined before trial heats were carried out. These were as follows:

• Initial (FetO+ MnO) % of ladle slag, that increases with increasing amount of
carry-over slag.
• Slag reducing agent amount (Al-dross).
• Reaction duration (ladle treatment duration which was until ladle were sent to
caster).
• Similar slag analysis data collected without using Al-dross; this was
necessary to see the difference when Al-dross was not used.

After the effect of these parameters were determined in this way, the relation of
this slag reducing practice to the reoxidation during continuous casting process, was
clarified. Using the data obtained for reoxidation of aluminium in steel (soluble
aluminium fade) as function of reducible oxide ( FetO + MnO ) amount in ladle slag,
a correlation was obtained.

42
CHAPTER 4

RESULTS AND DISCUSSION

4.1 Basic Oxygen Furnace Slag Chemistry

In Table 4.1, basic oxygen furnace slag analyses of low carbon aluminium
killed (LCAK) steel grades are given. The CaO and SiO2 analyses data of Erdemir
converter slags given in this table are represented as block diagrams in Figure 4.1.
From this figure, within limits, the average values of %CaO and %SiO2 in the
converter slag were determined as 49.6 and 10.7, respectively. When the basicity (B=
%CaO/%SiO2) of slag in equilibrium with liquid steel is between 2 and 4, it is known
that the ratio of %Fe2O3/%FeO is less than 0.05 [2,3]. However, calculated average
basicity of Erdemir converter slag was 4.6 and so it was out of the range 2-4;
therefore, the ratio of %Fe2O3/%FeO was not known exactly for this basicity. In any
case, the ratio of 4.6 was not much out of the given range, so, iron oxide was taken
entirely as FeO and FetO + MnO was used to represent the total reducible oxides. Of
course, metallic iron is excluded from the total iron analysed by X-Ray Fluorescence.

Using the data given in Table 4.1, FetO and MnO contents of Erdemir converter
slags are drawn as block diagrams in Figure 4.2. It was seen that average percentages
of FetO and MnO in converter slags were 26.66 and 3.76, respectively.

43
Table 4.1. Analyses of Basic Oxygen Furnace Slag for LCAK Steels (%)

No
of FetO SiO2 MnO Al2O3 CaO MgO P2O5 S Na2O K2O TiO2 Cr2O3
Heat
1 29.34 9.72 4.73 1.29 45.44 7.32 1.407 0.048 0.001 0.008 0.271 0.273
2 25.34 10.52 3.66 0.57 54.08 2.38 1.448 0.016 0.001 0.003 0.253 0.132
3 23.60 10.63 3.68 0.72 53.48 6.94 1.214 0.024 0.001 0.018 0.249 0.129
4 27.20 10.23 3.34 0.92 50.87 6.11 1.174 0.386 0.001 0.015 0.129 0.148
5 39.92 6.99 2.52 0.99 44.65 4.43 0.835 0.076 0.001 0.012 0.154 0.156
6 25.33 11.37 3.90 1.42 49.06 8.18 1.306 0.036 0.001 0.020 0.350 0.127
7 26.49 8.36 3.81 0.83 54.73 2.65 1.305 0.013 0.001 0.008 0.242 0.125
8 27.76 11.06 3.02 1.28 49.58 6.87 1.308 0.095 0.001 0.011 0.145 0.141
9 28.88 10.49 4.01 0.90 48.34 6.37 1.100 0.082 0.001 0.014 0.336 0.135
10 25.54 12.06 3.68 0.94 49.92 7.61 1.324 0.112 0.001 0.010 0.245 0.146
11 31.58 11.31 5.35 2.09 42.08 6.43 1.311 0.301 0.002 0.020 0.365 0.244
12 27.62 12.18 3.86 1.30 48.77 3.79 1.217 0.055 0.001 0.007 0.312 0.206
13 29.32 9.61 4.07 0.96 47.13 8.00 1.266 0.029 0.001 0.013 0.290 0.152
14 24.81 11.52 3.21 1.05 51.05 7.50 1.411 0.070 0.001 0.010 0.239 0.181
15 23.26 9.48 3.73 0.67 57.13 2.66 1.642 0.014 0.001 0.004 0.219 0.144
16 21.70 10.71 3.59 1.09 54.70 6.99 1.290 0.040 0.001 0.009 0.314 0.093
17 30.17 9.89 4.76 1.05 44.73 8.01 1.676 0.025 0.001 0.012 0.194 0.193
18 26.88 7.55 3.47 0.67 53.80 5.82 1.329 0.044 0.001 0.007 0.196 0.137
19 27.32 9.61 3.33 2.35 46.68 7.98 0.825 0.353 0.001 0.033 0.216 0.071
20 31.08 10.95 3.39 0.48 47.91 4.36 1.188 0.041 0.001 0.008 0.325 0.220
21 25.81 11.89 3.46 0.90 52.30 2.16 1.367 0.025 0.001 0.006 0.294 0.142
22 24.12 9.89 3.61 0.56 56.30 1.40 1.596 0.001 0.001 0.005 0.234 0.122
23 22.41 13.88 3.18 0.66 50.87 7.72 1.432 0.095 0.001 0.012 0.272 0.107
24 25.29 12.17 3.59 1.12 48.25 7.93 1.347 0.016 0.001 0.016 0.239 0.124
25 22.70 10.67 3.05 1.56 56.37 6.26 1.220 0.048 0.001 0.034 0.165 0.081
26 26.05 8.73 4.50 0.91 51.06 6.03 1.432 0.051 0.001 0.013 0.211 0.152
27 27.69 12.17 4.22 1.04 46.06 8.06 1.386 0.039 0.001 0.012 0.254 0.172
28 26.34 10.80 2.89 1.24 50.91 6.41 1.278 0.075 0.001 0.007 0.180 0.141
29 36.56 7.75 4.28 0.97 42.40 6.40 1.362 0.060 0.001 0.008 0.259 0.180
30 26.49 7.95 4.00 1.08 51.70 0.89 1.116 0.062 0.001 0.009 0.180 0.137
31 25.60 9.68 4.03 3.30 45.52 9.98 0.921 0.282 0.001 0.048 0.179 0.102
32 23.29 11.78 3.10 0.58 42.75 4.22 1.022 0.018 0.001 0.008 0.235 0.161
33 31.03 10.50 4.06 1.03 45.28 6.96 1.465 0.032 0.001 0.019 0.236 0.174
34 22.84 13.02 3.80 1.83 49.89 7.40 1.407 0.073 0.001 0.012 0.348 0.200
35 24.95 11.42 5.67 1.03 46.20 8.66 1.761 0.004 0.001 0.010 0.386 0.159
36 20.83 10.83 2.93 0.77 57.94 5.83 1.247 0.023 0.001 0.014 0.191 0.112
37 27.07 12.09 3.32 0.90 48.50 6.18 1.351 0.033 0.001 0.010 0.195 0.123
38 25.96 12.56 4.50 0.99 46.32 8.53 1.368 0.024 0.001 0.016 0.417 0.170
39 32.47 8.76 3.73 0.94 44.97 7.02 1.399 0.058 0.001 0.013 0.203 0.230
40 20.74 13.01 3.55 1.12 55.45 2.48 1.565 0.012 0.001 0.006 0.298 0.107
41 23.00 13.33 3.85 1.75 51.05 3.55 1.416 0.032 0.001 0.008 0.299 0.145
42 27.82 13.04 3.12 1.82 48.08 4.67 1.132 0.062 0.001 0.023 0.263 0.333

44
Table 4.1(Continued)Analyses of Basic Oxygen Furnace Slag for LCAK Steels (%)

No
of FetO SiO2 MnO Al2O3 CaO MgO P2O5 S Na2O K2O TiO2 Cr2O3
Heat
43 27.33 9.46 4.67 1.26 48.13 8.99 1.169 0.012 0.001 0.013 0.230 0.185
44 28.22 10.41 3.67 1.11 47.89 7.44 1.314 0.044 0.001 0.012 0.213 0.174
45 29.72 8.09 3.25 1.27 49.91 5.83 0.986 0.073 0.001 0.010 0.240 0.167
46 20.48 14.73 4.45 1.50 52.17 3.49 1.670 0.004 0.001 0.008 0.415 0.128
47 23.06 12.08 3.94 1.41 53.09 3.08 1.327 0.102 0.001 0.010 0.363 0.198
48 27.62 10.09 3.79 1.52 47.52 8.17 1.342 0.041 0.001 0.009 0.232 0.208
49 24.83 11.20 3.49 1.22 50.82 7.78 1.155 0.042 0.001 0.013 0.272 0.156
50 30.05 8.61 3.37 1.02 48.09 6.82 1.055 0.095 0.001 0.009 0.255 0.142
Avg. 26.67 10.70 3.76 1.16 49.60 6.01 1.30 0.07 0.00 0.01 0.26 0.16

30 30

25 25

20 20
% of Heats

% of Heats

15 15

10 10

5 5

0 0
42 44 46 48 50 52 54 56 58 Diğer 6 7 8 9 10 11 12 13 14 Diğer
% CaO % SiO2

Figure 4.1 CaO and SiO2 contents of Erdemir BOF Slag

45
25 30

25
20

20
15

% of Heats
% of Heats
15
10

10
5
5

0
0
2.5 2.8 3.1 3.4 3.7 4.0 4.3 4.6 4.9 Diğer
17 19 21 23 25 27 29 31 33 35 Diğer
%MnO % FetO

Figure 4.2 FetO and MnO contents of Erdemir BOF Slag

4.2 Carry-over Slag Weight

After the primary oxygen steelmaking process, when ladle was ready for
secondary treatment process, slag contained 5 main ingredients. These were:

• Burnt Lime as flux ( 98 % CaO): 600 kg. for every heat.


• Al2O3 formed during tapping by aluminium bar addition.
• Ladle slag remained from previous heat.
• Carry-over slag from converter.
• Eroded refractory from ladle walls.

To find the quantity of carry-over slag following equation was written:

Wst = Wf + WsAl2O3 + Wsl + Wsc + Wsr (46)

where,

Wst : total weight of slag, kg


Wf : weight of flux (burnt lime) added during tapping, kg .
WsAl2O3 : weight of alumina formed after deoxidation, kg .

46
Ws l : weight of slag remained from previous heat, kg .
Ws c : weight of carry-over slag tapped from converter, kg .
Ws r : weight of refractory eroded from ladle lining, kg .
Then,
Wsc = Wst - Wf - WsAl2O3 - Wsl - Wsr (47 )

In order to find the total weight of slag (Wst ), below formula for initial alumina
content of slag after tapping steel from converter, was used:

%(Al2O3)i= [ [WsAl2O3] + ( Wsl * (%Al2O3)a /100)] / Wst (48)

where,

WsAl2O3 = [(WAlt – WAla) * MAl2O3 / 2MAl ]


and,
WAlt : weight of aluminium bar added during tapping from converter, kg .
WAla : weight of aluminium that dissolved in liquid steel (from initial steel
analysis at the ladle furnace)
MAl2O3 : molecular weight of Al2O3
MAl : atomic weight of Al
(%Al2O3)a : average final (Al2O3) content in ladle slag after ladle treatment.

(%Al2O3)a was taken as 26%, which was the average of experimental results
given in Table 4.2, and since Wsl can not be calculated, this value was taken as 100
kg . Also, since the weight difference between new-lined ladle and a ladle that made
75 heats was 7500 kg (average value), Wsr was taken as 100 kg for every heat. The
results of calculations done using the formula given above are given in Table 4.2.

47
Table 4.2. Calculated Weight of Carry-over Slag Using Experimental Data

a
WAl
WAlt (kg) % (Al2O3)i %Al WsAl2O3 (kg) Wst (kg) Wsc (kg)
(kg)
175 16.49 0.062 74 319 2091 946
220 21.15 0.017 20 412 2073 834
200 21.53 0.032 38 372 1847 650
220 23.04 0.033 40 409 1889 654
215 19.5 0.045 54 398 2172 949
210 18.86 0.053 64 387 2188 975
210 20.65 0.068 82 384 1985 775
200 18.91 0.056 67 367 2079 886
200 23.41 0.022 26 374 1707 507
230 27.26 0.002 2 434 1688 428
220 26.75 0.011 13 413 1643 403
180 18.77 0.036 43 333 1914 754
200 24.51 0.045 54 369 1613 417
225 21.01 0.032 38 419 2118 873
220 28.52 0.016 19 413 1538 299
200 22.34 0.024 29 373 1787 588
220 22.81 0.026 31 411 1914 678
245 25.28 0.011 13 461 1925 639
175 24.16 0.029 35 325 1453 302
210 27.97 0.05 60 387 1477 264
240 28.59 0.004 5 453 1674 395
230 18.86 0.066 79 422 2375 1127
200 25.55 0.041 49 370 1550 354
210 25.58 0.058 70 386 1610 398
190 19.82 0.028 34 354 1915 736
200 22.47 0.026 31 373 1775 576
200 20.28 0.029 35 372 1964 766
235 27.93 0.02 24 440 1669 403
240 23.99 0.053 64 443 1956 687
230 25.11 0.017 20 431 1821 564
235 30.23 0.016 19 441 1544 278
220 22.06 0.043 52 407 1965 731
180 20.02 0.005 6 339 1823 658
200 24.79 0.027 32 373 1608 410
220 23.57 0.036 43 409 1845 610
200 25.74 0.009 11 376 1562 360
210 20.85 0.033 40 390 1997 781
210 21.31 0.045 54 388 1944 729
170 15.52 0.025 30 316 2206 1064
220 19.72 0.04 48 408 2201 967

48
Table 4.2.(Continued) Calculated Weight of Carry-over Slag Using Experimental
Data

a
WAl
WAlt (kg) % (Al2O3)i %Al WsAl2O3 (kg) Wst (kg) Wsc (kg)
(kg)
220 18.17 0.048 58 406 2380 1148
225 29.14 0.032 38 419 1527 282
210 23.22 0.06 72 385 1771 560
230 28.3 0.015 18 432 1617 359
230 20.99 0.02 24 431 2176 919
220 24.78 0.045 54 407 1748 515
220 20.24 0.037 44 409 2147 913
230 21.16 0.022 26 430 2156 900
240 28.05 0.003 4 453 1707 428
230 21.69 0.068 82 422 2064 816
200 24.23 0.022 26 374 1649 450
220 22.04 0.033 40 409 1975 740
240 27.48 0.038 46 446 1718 446
200 23.21 0.024 29 373 1720 521
200 22.45 0.028 34 372 1775 577
220 21.67 0.041 49 408 2002 768
200 18.97 0.038 46 371 2091 894
200 20.94 0.078 94 363 1858 669
190 20.52 0.018 22 355 1859 678
200 23.04 0.033 40 372 1725 528
200 27.19 0.008 10 376 1479 277
210 25.13 0.023 28 392 1665 446
210 20.11 0.042 50 389 2062 848
210 19.33 0.047 56 388 2141 927
220 18.67 0.049 59 406 2315 1083
220 23.74 0.051 61 406 1819 587
225 21.77 0.032 38 419 2044 799
225 25.08 0.035 42 418 1772 527
200 22.69 0.038 46 371 1748 551
200 17.02 0.057 68 367 2309 1116
220 21.8 0.023 28 411 2006 768
220 20.52 0.032 38 410 2122 887
180 24.08 0.024 29 335 1501 340
212 23 0.034 40 394 1873 653

49
As it can be seen from Table 4.2 carry-over slag amount changed between 264 kg
and 1148 kg. The possible reasons for this can be listed as:

• Erosion in tapping hole of the converter


• Variation in converter slag viscosity
• Difference between operator experiences.

4.3 Effect of Initial Slag Composition on Final Reducible Oxide Percentage

In this part of the thesis work, it was aimed to assess the data according to
relationship of initial and final reducible oxide amount of ladle slag. Since the weight
and composition of the carry-over slag would change, initial total reducible oxide
amount of ladle slag was between 7-19 % and on the average it was taken as 13.49
%. From a standard ladle treatment process that consisted 100 kg. of aluminium
dross addition and 45 minutes of average reaction time, final reducible oxide
(R.O.)% in slag was analysed.

50
Table 4.3 Ladle Furnace Treatment Data Including Initial and Final Slag
Compositions

Data Steel (FetO+MnOi (FetO+MnO)f Reaction Al Dross Wsc Wst


No Grade (%) (%) Time (min) Amount (kg) (kg.) (kg.)
1 7110K 8.65 6.23 42 100 547 1554
2 7112K 9.93 6.41 61 100 491 1503
3 7110K 10.31 4.12 33 100 692 1643
4 7112K 10.25 4.29 50 100 488 1470
5 7110K 10.65 6.89 66 100 433 1427
6 7112K 11.36 6.03 49 100 505 1516
7 7112K 12.10 6.96 40 100 410 1378
8 7112K 12.19 5.64 31 100 400 1366
9 7112K 12.33 4.27 44 100 410 1369
10 7112K 12.66 5.14 58 100 633 1604
11 7112K 12.65 5.55 44 100 684 1673
12 7112K 12.31 4.56 47 100 534 1508
13 7112K 13.25 4.52 43 100 679 1660
14 7112K 13.70 5.96 48 100 704 1672
15 7112K 13.91 10.02 50 100 588 1559
16 7112K 14.08 7.89 43 100 645 1681
17 7112K 14.17 8.10 37 100 664 1623
18 7112K 14.57 9.72 51 100 715 1669
19 7112K 15.07 8.09 35 100 459 1408
20 7112K 15.22 9.90 41 100 483 1457
21 7112K 15.44 7.60 68 100 810 1797
22 7112K 15.43 8.94 33 100 641 1669
23 7112K 15.42 8.76 34 100 787 1762
24 7112K 16.36 10.01 35 100 530 1501
25 7112K 16.57 12.69 46 100 262 1213
26 7112K 16.50 8.74 33 100 804 1765
27 7112K 19.19 11.07 65 100 677 1630
Avg. 13.49 7.34 45 100 580 1558

51
It was seen from data given in Table 4.3 that the initial reducible oxide (FetO +
MnO)1 % in slag varied between 8-19 %, and, as initial reducible oxide percentage
in slag decreased, final reducible oxide (FetO + MnO)f % decreased also, which
varied between 4-12 %. This relation is given graphically in Figure 4.3.

14

12

10
Final (FetO+MnO)%

0
8 9 10 11 12 13 14 15 16 17 18 19 20

Initial (FetO+MnO)%

Figure 4.3. Effect of Initial Reducible Oxide Content on Final Reducible Oxide %

For initial (FetO + MnO) % =14, it decreases to 8% with 100 kg of Al dross


addition. If average total ladle slag weight is taken as 1558 kg, total weight of
reduced FeO and MnO is calculated as:

WRFeO+MnO = ∆(FeO + MnO) * Wst / 100 = 93.4 kg


WAl = WRFeO+MnO *2 MAl / (3 * MR.O.) = 23.5 kg

WRFeO+MnO = weight of (FeO + MnO) reduced after ladle treatment, kg


∆(FeO + MnO) = difference in (FeO + MnO) after ladle treatment, %
WAl = weight of aluminium used for reduction, kg
MR.O. = average molecular weight of R.O., kg

52
4.4. Effect of Reaction Duration on Final Reducible Oxide Percentage

In order to analyse the process kinetically, the treatment time at the ladle
furnace was taken as a variable parameter and called ‘reaction duration’. Other
parameters such as initial R.O. %in slag (data taken in a narrow range of 11-14 %)
and Al dross addition (100 kg) were taken constant.

The data given in Table 4.4 were used to determine the effect of reaction duration
on the final reducible oxide percentage of ladle treatment slag. It was clear from
these data that the longer the reaction duration, lower would be the final reducible
oxide content of slag.

Table. 4.4. Data to Show the Effect of Reaction Duration

Data Steel (FetO+MnO)i (FetO+MnO)f Reaction Al Dross


Wsc (kg.) Wst (kg.)
No Grade % % Time (min) (kg)
1 7112K 13.87 9.63 27 100 463 1424
2 7112K 11.59 9.70 27 100 517 1493
3 7112K 14.38 10.22 28 100 606 1641
4 7112K 12.19 9.47 30 100 544 1500
5 7112K 12.19 5.64 31 100 400 1366
6 7112K 14.17 8.10 37 100 664 1623
7 7112K 12.11 7.12 40 100 410 1378
8 7112K 13.25 7.93 43 100 679 1660
9 7112K 14.08 7.89 43 100 645 1681
10 7112K 12.32 5.01 44 100 410 1369
11 7112K 12.65 5.55 44 100 684 1673
12 7112K 12.31 4.56 47 100 534 1508
13 7112K 13.70 5.96 48 100 704 1672
14 7112K 11.28 6.03 49 100 505 1516
15 7112K 12.66 5.14 58 100 633 1604
Avg. 12.85 7.20 40 100 560 1540

53
The data in Table 4.4 was used to plot Figure 4.4. It was seen from this figure that
for an average ladle treatment duration of 30 minutes, the final FetO + MnO % in
slag was between 8 and 9, and as the duration increased to 45-50 min. the reducible
oxide content of final slag decreased to 4 to 5 %.

12

10
Final (FetO+MnO)%

0
25 30 35 40 45 50 55 60
Reaction Tim e (m in.)

Figure 4.4 Effect Reaction Time on Final Reducible Oxide %

4.5. Effect of Aluminium Dross Amount on the Final Reducible Oxide


Percentage

To understand the relation between amount of aluminium dross used and final
reducible oxide content of slag after ladle treatment, similar trial heats were made
with 60, 80, 100, 120 and 150 kg aluminium dross addition during ladle treatment.
During these trials data taken constant were initial FetO + MnO content of slag as
around 12.5%, average dissolved aluminium in steel as 0.035% and average reaction
duration as 35 minutes. Then, decrease in the final reducible oxide in slag was

54
calculated from the analysis of slag for R.O. %. The data are shown in Table 4.5 and
are given graphically in Figure 4.5.

Table 4.5 Data to Show The Effect of Al Dross Amount (kg)

Al Dross (Fe O+MnO)


Heat No Steel Grade (FetO+MnO)i(%) (FetO+MnO)f (%) ∆ t
(kg) (%)
1 60 7112K 11.99 8.08 3.91
2 80 7112K 13.09 7.80 5.29
3 100 7112K 12.74 6.44 6.30
4 120 7112K 13.52 7.03 6.49
5 150 7112K 11.57 3.71 7.86

7
Final (FetO + MnO)%

2
50 70 90 110 130 150 170
Al Dross (kg)

Figure 4.5. Effect of Aluminium Dross Addition to Ladle Treatment Slag.

55
4.6 Slag Reduction by Dissolved Aluminium in Steel

As it was indicated before, lowering the oxidation potential of ladle slag was a
reduction reaction between aluminium, and, iron and manganese oxides as given by
reactions (49) and (50) below.

3(FetO) + 2[Al] = (Al2O3) + 3[Fe] (49)

3(MnO) + 2[Al] = (Al2O3)+ 3[Mn] (50)

Therefore, dissolved aluminium in liquid steel would also reduce the reducible
oxides in slag and in order to clarify the effect of aluminium dross usage, ten extra
heats were produced without aluminium dross addition. The results are given in
Table 4.6.

Table 4.6. Data for Heats With and Without Aluminium Dross Addition

Data Steel (FetO+MnO)i (FetO+MnO)f Reaction Al Dross Wsc Wst


No Grade (%) (%) Time (min) (kg) (kg.) (kg.)
1 7112K 10.31 4.12 33 100 692 1643
2 7112K 10.76 5.40 39 100 799 1831
3 7112K 12.19 5.64 31 100 400 1366
4 7112K 12.59 4.52 26 100 309 1295
5 7112K 13.87 9.63 27 100 463 1424
6 7112K 14.17 7.96 37 100 664 1623
7 7112K 14.38 6.61 28 100 606 1641
8 7112K 15.07 8.09 35 100 459 1408
9 7112K 15.43 8.94 33 100 641 1669
10 7112K 16.36 10.01 35 100 530 1501
11 7112K 9.74 7.87 24 - 305 1384
12 7112K 13.59 9.23 36 - 385 1443
13 7112K 11.02 7.55 21 - 827 1864
14 7112K 17.81 12.12 30 - 761 1848
15 7112K 12.46 7.81 25 - 620 1762
16 7112K 9.41 6.03 32 - 595 1654
17 7112K 14.03 10.04 29 - 534 1596
18 7112K 15.98 9.94 35 - 642 1632
19 7112K 8.70 5.90 36 - 429 1495
20 7112K 16.44 12.60 27 - 565 1674

56
For a fixed reaction time (average 31 min.) initial and final slag compositions are
obtained and differences between two cases, with and without Al-dross addition,
can be seen from Figure 4.6.

14

12

10 (a)
Final (FetO+MnO)%

6
(b)
4

0
8 10 12 14 16 18 20
Initial (Fe tO+MnO)%

Figure 4.6 Reduction (a) without and (b) with Al Dross Addition

For an initial reducible oxide content of about 10-12 %, using 100 kg


aluminium dross, after about 30 minutes of ladle treatment, it decreased to 4.5-5.0 %
R.O. On the other hand, without using the aluminium dross, the reducible oxide
content of slag decreased from 12% R.O. to about 7.5 %. This meant that dissolved
aluminium in steel did some reduction but it was more effective by aluminium dross
usage.

57
4.7. Effect of Final Reducible Oxide Content on Reoxidation of Steel

As explained in section 2.4.4.4, FetO + MnO represented the oxygen potential


of slag and acted as reoxidation source for steel. To be able to see the effect of
reducible oxide content of slag on reoxidation of steel, another steel grade (ultra low
carbon) produced with a different process route was selected and compared with
typical low carbon aluminium killed grade, according to aluminium dissolved(loss)
during continuous casting. Table 4.7 gives comparison of these two grades.

Table 4.7. Comparison of Low Carbon and Ultra Low Carbon Grades of Steel

ULC GRADES LC GRADES


PLANT RH-KTB Ladle Furnace
%C 0.003-0.004 0.02-0.04
% Final R.O. 16-17 6-7
%Al 0.02-0.06 0.02-0.06
Tundish Cover Ca-aluminate + rice hulls Ca-aluminate + rice hulls
Other Casting Parameter Same Same

For ultra low carbon (ULC) aluminium killed steel production, RH-KTB type
vacuum degasser is used to reach lower carbon than basic oxygen furnace and
deoxidation of steel was done after vacuum treatment. It meant that reduction of slag
could not be carried out because of process type and deoxidation method.

58
In RH-KTB plant steel was just circulating (60 ton/min), so slag-metal
reaction was not so effective in this process. Also, deoxidation was carried out at the
end of the vacuum degassing process rather than in BOF and naturally final average
reducible oxide percentage was 16-17% while it was 6-7 % for low carbon
aluminium killed steel grade which was treated with aluminium dross and which is
the main topic of this thesis.

In order to see the effect of difference in between reducible oxide contents of


these two different steel groups, reoxidation (or dissolved aluminium percentage in
steel) during casting because of air entry from nozzles, and, oxygen sources as
tundish flux and ladle slag was analysed for 1000 heats from each group. Figures 4.7
and 4.8 show the distribution of dissolved aluminium loss in Ultra-Low Carbon
(ULC)and Low Carbon (LC) steels, respectively

14.0

12.0

10.0

8.0
% of Heats

6.0

4.0

2.0

0.0
3

7
0

00

00

00

01

01

01

02

02

02
0.

0.

0.

0.

0.

0.

0.

0.

0.

Al Loss (%) During Casting

Figure 4.7. Aluminium Loss Distribution during Casting Ultra-Low Carbon Steel
Grades.

59
16.0

14.0

12.0

% of Heats 10.0

8.0

6.0

4.0

2.0

0.0
3

7
0

00

00

00

01

01

01

02

02

02
0.

0.

0.

0.

0.

0.

0.

0.

0.
Al Loss (%) During Casting

Figure 4.8. Aluminium Loss Distribution during Casting Low Carbon Steel Grades.

As it can be seen from Figures 4.7 and 4.8 average aluminium losses for
ULC and LC aluminium killed steel grades were, on the average, 144 ppm. and 94
ppm., respectively. As a result, alumina content of ULC steel was being expected to
be more than LC grade steel and this would affect the steel internal cleanliness,
mechanical properties, and would increase the nozzle blockage frequency.

60
CHAPTER 5

CONCLUSIONS

The aim of this study was to investigate how the reducibility of Erdemir ladle
slag changed according to certain properties by using aluminium dross as slag
refiner. For this study Erdemir Steelmaking Plant Basic Oxygen Furnace and Ladle
Furnace slag analysis were used. This slag was produced during production of low
carbon steel containing 0.02-0.04 %C, 0.10-0.20 %Mn, 0.020 %P (max.), 0.020 %S
(max.) and 0.020-0.060 %Al.

Slag samples, collected with a steel rod from slag-steel contact surface, were
analysed by XRF to determine slag composition. Converter (BOF) slags were mainly
composed of CaO, FetO, SiO2, MgO, MnO and P2O5, and, ladle slags were
composed of CaO, Al2O3, SiO2, MgO, FetO and MnO.

For aluminium killed steel grades FetO and MnO in slag represented the
oxidizing capacity of ladle slag and were carried with carry-over slag of converter
which leaked to ladle during tapping. The higher the FetO and MnO contents (called
R.O.- reducible oxides) in the ladle slag, the greater was the potential for reoxidation
and the corresponding generation of alumina inclusions. Many slivers in the final
product were traced to reoxidation that originated from FetO and MnO in the ladle
slag.

There were several factors which affected the reducing ability of aluminium
dross that was tried to be used in the ladle furnace in Erdemir and these were, amount
of aluminium dross, initial reducible oxide content of steelmaking slag, reaction
(treatment) duration and reduction with dissolved aluminium in steel.

61
The amount of reducible oxides in the mentioned steelmaking slags decreased
with increase in the amount of aluminium dross added. For an average of 12-13 %
initial reducible oxide in slag, addition of 100 kg. aluminium dross decreased final
reducible oxide in slag to 7-8 %. As the amount of initial reducible oxide content of
slags increased, the final reducible oxide content of slags also increased, but the
degree of reducible oxide content decrease was relatively high when the treatment
was started with high reducible oxide in slag. In any case a low initial reducible
content obtained during converting of steel would be preferred.

As the treatment duration in the ladle furnace increased, reaction of aluminium


dross with the slag increased and so reducible oxide content of slag decreased.
Comparing the heats with and without the use of aluminium dross (dissolved
aluminium in steel acted as deoxidizer) showed that aluminium dross usage resulted
in comperatively more decrease in reducible oxide content of ladle furnace slags.

Reoxidation of steel shown by the soluble aluminium loss (fade) during casting
was higher in the ultra-low carbon steels which were not treated with aluminium
dross, as compared to low-carbon aluminium killed steels which were the types
considered in this work. This oxidation would promote ladle and tundish nozzle
clogging, and casting problems while decreasing steel cleanliness.

62
REFERENCES

[1] Andersson M., Hallberg M., Jonsson L., and Jonsson P., “Slag–Metal Reactions
during Ladle Treatment with Focus on Desulphurisation”, Ironmaking and
Steelmaking, 29(3), p. 224-232, 2002.

[2] Huh W. W., Jung W-G., “Effect of Slag Composition on Reoxidation of


Aluminium Killed Steel”, ISIJ International, 36(9),p136-139, 1996.

[3] Turkdogan E.T., “Rationale on Composition of Slags in Oxygen Steelmaking


Processes”, Iron and Steelmaking, 26(5), p.358-362, 1999.

[4] Beskow K., Sichen Du., “Ladle Glaze: Major Source of Oxide Inclusions during
Ladle Treatment of Steel”, Ironmaking and Steelmaking 31(5), p.393-399, 2004.

[5] Peter J., Peaslee K.D., Robertson D.G.C., Thomas B.G.,“Experimental Study of
Kinetic Processes during the Steel Treatment at Two LMF’s”, AISTech 2005
Proceedings, Vol.1, p.959-973, 2005.

[6] Conejo A.N., Hernandez D.E., “Analysis of Aluminium Deoxidation Practice in


the Ladle Furnace”, AISTech 2005 Proceedings, Vol.l, p.947-957, 2005

[7] Millman M.S., “Clean Steel Fundamentals”, International Iron and Steel Institute,
p 1-6, 2005.

[8] Zhang L., Thomas B.G., “State of the Art in Evaluation and Control of Steel
Cleanliness” . ISIJ 43 (3), p.271-291, 2003.

[9] Turkdogan E.T., “Fundamentals of Steelmaking”, The Institute of


Materials,p.180-259, 1996.

63
[10] Lankford W.T., Samways N.L., Craven R.F., McGannon H.E.,“The Making,
Shaping and Treating of Steel”, United States Steel, 1982.

[11] Sumida M., Nadif M., Burty M., Tusset V., “ULC and LC Steel Qualities”, IISI
p.365-406, 2005.

[12] Nadif M., Burty M., Soulard H., Boher M., Pusse C., Lehmann J., Ruby-Meyer
F., Guiban M.A., “Reoxidation in Steel and CC Nozzle Blockage”, IISI, p.87-152,
2005.

[13] Fruehan R.J., “Ladle Metallurgy Principles and Practices”, Iron and Steel
Society, p.21-42, 1985.

[14] Lucheva B., Tosenov Ts., Petkov R., “Non-waste Aluminium Dross Recycling”
, Journal of the University of Chemical Technology and Metallurgy, p 335-338, 2005

[15] Ghorab H.Y., Rizk M., Matter A., Salama A.A., “Characterization and
Recycling of Aluminium Slag”, Polymer-Plastics Technology and Engineering,
p.1663-1673, 2004

64

You might also like